首页 | 本学科首页   官方微博 | 高级检索  
相似文献
 共查询到20条相似文献,搜索用时 15 毫秒
1.
The UV RESISTANCE LOCUS 8 (UVR8) photoreceptor specifically mediates photomorphogenic responses to UV‐B. Photoreception induces dissociation of dimeric UVR8 into monomers to initiate responses. However, the regulation of dimer/monomer status in plants growing under photoperiodic conditions has not been examined. Here we show that UVR8 establishes a dimer/monomer photo‐equilibrium in plants growing in diurnal photoperiods in both controlled environments and natural daylight. The photo‐equilibrium is determined by the relative rates of photoreception and dark‐reversion to the dimer. Experiments with mutants in REPRESSOR OF UV‐B PHOTOMORPHOGENESIS 1 (RUP1) and RUP2 show that these proteins are crucial in regulating the photo‐equilibrium because they promote reversion to the dimer. In plants growing in daylight, the UVR8 photo‐equilibrium is most strongly correlated with low ambient fluence rates of UV‐B (up to 1.5 μmol m?2 s?1), rather than higher fluence rates or the amount of photosynthetically active radiation. In addition, the rate of reversion of monomer to dimer is reduced at lower temperatures, promoting an increase in the relative level of monomer at approximately 8–10 °C. Thus, UVR8 does not behave like a simple UV‐B switch under photoperiodic growth conditions but establishes a dimer/monomer photo‐equilibrium that is regulated by UV‐B and also influenced by temperature.  相似文献   

2.
Insulin-receptor interactions in liver cell membranes   总被引:17,自引:0,他引:17  
The specific binding of 125I-insulin to liver cell membranes is a saturable process with respect to insulin. Binding is displaced by low concentrations of native insulin but not by biologically inactive insulin derivatives or by other peptide hormones. The rate constants of association (3.5 × 106 mole−1 sec−1) and of dissociation (2.7 × 10−4 sec−1) of the insulin-membrane complex can be determined independently. The dissociation constant of the complex, determined from the rate constants and from equilibrium data, is about 7 × 10−11M. Complex formation does not result in degradation of the insulin molecule. The binding interaction is a dissociable process involving a homogeneous membrane structure which is almost certainly the biologically significant receptor. The kinetic properties, and the effects of enzymic perturbations of the membrane, suggest that the insulin receptors of liver and of adipose tissue cells may be very similar structures.  相似文献   

3.
The finding that exchange of tubulin subunits between tubulin dimers (alpha-beta + alpha'beta' <--> alpha'beta + alphabeta') does not occur in the absence of protein cofactors and GTP hydrolysis conflicts with the assumption that pure tubulin dimer and monomer are in rapid equilibrium. This assumption underlies the many physical chemical measurements of the K(d) for dimer dissociation. To resolve this discrepancy we used surface plasmon resonance to determine the rate constant for dimer dissociation. The half-time for dissociation was approximately 9.6 h with tubulin-GTP, 2.4 h with tubulin-GDP, and 1.3 h in the absence of nucleotide. A Kd equal to 10(-11) M was calculated from the measured rate for dissociation and an estimated rate for association. Dimer dissociation was found to be reversible, and dimer formation does not require GTP hydrolysis or folding information from protein cofactors, because 0.2 microM tubulin-GDP incubated for 20 h was eluted as dimer when analyzed by size exclusion chromatography. Because 20 h corresponds to eight half-times for dissociation, only monomer would be present if dissociation were an irreversible reaction and if dimer formation required GTP or protein cofactors. Additional evidence for a 10(-11) M K(d) was obtained from gel exclusion chromatography studies of 0.02-2 nM tubulin-GDP. The slow dissociation of the tubulin dimer suggests that protein tubulin cofactors function to catalyze dimer dissociation, rather than dimer assembly. Assuming N-site-GTP dissociation is from monomer, our results agree with the 16-h half-time for N-site GTP in vitro and 33 h half-life for tubulin N-site-GTP in CHO cells.  相似文献   

4.
Two processes of porphyrins in heterogeneous systems containing aqueous and membrane phases have been studied with hematoporphyrin and hematoporphyrin derivative: Dimerization equilibrium in the aqueous phases and porphyrin-membrane binding equilibrium using liposomes as models for biological membranes. The interrelationship of aqueous aggregations and membrane binding was probed and the porphyrin aggregation state in the membrane, at equilibrium, was assessed. Fluorimetric techniques were employed. The dimerization equilibrium constants, at neutral pH and 37°C were found to be 2.8 · 105 M−1 and 1.9 · 106 M−1 for hematoporphyrin and its derivative, respectively. Over a porphyrin concentration range going from monomer-dominant to dimer-dominant systems, we have found that only monomers are bound to the membrane. The respective monomer-liposome binding constants, found to be independent of the initial monomer/dimer distribution in the aqueous phase, were determined to be 1.6 · 103 M−1 and 4.1 · 103 M−1 at neutral pH and 37°C for hematoporphyrin and its derivative, respectively. The monomer-liposome interaction was found to perurb the initial monomer/dimer distribution in the aqueous phase, so that the monomers residing at equilibrium in the membrane originate from both monomers and dimers in the aqueous phase.  相似文献   

5.
The enzymatic activity of d-glyceraldehyde 3-phosphate dehydrogenase depends nonlinearly on protein concentration in the range 3 × 10?8 to 3 × 10?6m. With increasing enzyme concentrations the apparently hyperbolic substrate saturation curves turn into sigmoidal ones. From the kinetic and physicochemical data it is assumed that the enzyme exists as an equilibrium mixture of different oligomeric states. The system is found to be consistent with a model characterized by rapid equilibrium between monomer-dimer-tetramer, the tetramer being inactive, assuming identical intrinsic binding constants for the substrate in the monomer and in the dimer.  相似文献   

6.
An investigation was made of the role of calcium ions in the reversible stage of fibrin polymerization, using a direct and relatively simple approach. Purified fibrin monomer in solution (7.5 mg/ml) in 1.0 m NaBr (pH 5.3) was polymerized by raising the pH to 5.7–7.7 by the addition of aliquots of standard NaOH solution and the rate and total extent of proton release during polymerization were measured potentiometrically. In the presence of added CaCl2 (10−5-10−2m) the rate of proton release was increased and the clotting time was decreased. The profile of equilibrium proton release vs pH of polymerization was also shifted, the maximum being increased and occurring at a lower pH. Sedimentation velocity studies in the intermediate pH range (5.7–6.0) showed that the altered profile of equilibrium proton release was due to a broadening of the pH range of polymerization, and that polymerization remained reversible in the presence of CaCl2. At pH 5.3, where fibrin is essentially monomeric, addition of CaCl2 resulted in the release of protons and small increases in sedimentation coefficient and reduced viscosity. Under the same conditions, a similar release of protons was observed from fibrinogen, but there was no effect on its sedimentation coefficient. It was concluded that the proton release at pH 5.3 was due mainly to binding of calcium ions to fibrinogen and fibrin monomer. The effect of CaCl2 on the sedimentation coefficient of fibrin at pH 5.3 was found to decrease with decreasing protein concentration, indicating that it was the result of a small extent of polymerization, rather than a conformational change. Added MgCl2 had no effect on fibrin monomer at pH 5.3 and no significant effect on the rate or extent of proton release during polymerization at higher pH, indicating that there are specific binding sites for calcium ions in fibrinogen and fibrin. The observed effects of bound calcium ions on reversible fibrin polymerization are explained most simply in electrostatic terms.  相似文献   

7.
Cyanide binding to Chromatium vinosum ferricytochrome c′ has been studied to further investigate possible allosteric interactions between the subunits of this dimeric protein. Cyanide binding to C. vinosum cytochrome c′ appears to be cooperative. However, the cyanide binding reaction is unusual in that the overall affinity of cyanide increases as the concentration of cytochrome c′ decreases and that cyanide binding causes the ligated dimer to dissociate to monomers as shown by gel-filtration chromatography. Therefore, the cyanide binding properties of C. vinosum ferricytochrome c′ are complicated by a cyanide-linked dimer to monomer dissociation equilibrium of the complexed protein. The dimer to monomer dissociation constant is 20-fold smaller than that for CO linked dissociation constant of ferrocytochrome c′. Furthermore, the pH dependence of both the intrinsic equilibrium binding constant and the dimer to monomer equilibrium dissociation constant was investigated over the pH range of 7.0 to 9.2 to examine the effect of any ionizable groups. The equilibrium constants did not exhibit a significant pH dependence over this pH range.  相似文献   

8.
S. R. Martin 《Biopolymers》1980,19(3):713-721
The self-association of daunorubicin in aqueous solution has been examined using visible absorption, fluorescence, and CD measurements. Spectral changes in the concentration range 10?6 to 1.5 × 10?3M have been interpreted in terms of a monomer–dimer equilibrium for daunorubicin. The data have been analyzed using a nonlinear curve-fitting technique. The results obtained in this study differ markedly from previously published results, and possible reasons for this difference are discussed. The effect of solvent, temperature, and ionic strength on the dimerization process is investigated.  相似文献   

9.
Dimeric intercellular adhesion molecule-1 (ICAM-1) has been known to more efficiently mediate cell adhesion than monomeric ICAM-1. Here, we found that truncation of the intracellular domain of ICAM-1 significantly enhances surface dimerization based on the two criteria: 1) the binding degree of monomer-specific antibody CA-7 and 2) the ratio of dimer/monomer when a mutation (L42 → C42) was introduced in the interface of domain 1. Mutation analysis revealed that the positively charged amino acids, including very membrane-proximal 505R, are essential for maintaining the structural transition between the monomer and dimer. Despite a strong dimer presentation, the ICAM-1 mutants lacking an intracellular domain (IC1ΔCTD) or containing R to A substitution in position 505 (505R/A) supported a lower degree of cell adhesion than did wild-type ICAM-1. Collectively, these results demonstrate that the native structure of surface ICAM-1 is not a dimer, but is an intermediate monomer–dimer equilibrium structure by which the effectiveness of ICAM-1 can be fully achieved.  相似文献   

10.
The energy dependence of the chemical properties of cytochrome c oxidase   总被引:3,自引:0,他引:3  
The addition of ATP to intact mitochondria induces a high- to low-spin state transition in a heme of oxidized cytochrome oxidase. This transition is dependent on the phosphate potential with one-half effect requiring a phosphate potential approximately 2 × 103m−1. The data are consistent with a phosphate potential dependent equilibrium between two oxidized forms of cytochrome oxidase (one oxidase per ATP). The addition of ATP to mitochondria decreases the rate of reaction of cyanide with oxidized cytochrome oxidase by at least 103 and modifies both affinity and spectral change induced by adding cyanide to reduced cytochrome oxidase.  相似文献   

11.
Radiation chemical studies of thioesculetin (1), a thioketone derivative of coumarin, were performed by both pulse radiolysis technique and DFT calculations. Hydroxyl (?OH) radical reaction with 1 resulted transients absorbing at 320, 360 and 500?nm. To identify the nature of the transients, the reaction was studied with specific one-electron oxidant (N3?) radical, where 360?nm band was absent. The transient absorption at 500?nm was concentration-dependent. The overall impression for ?OH radical reaction was that the transient absorbing at 320, 360 and 500?nm was due to sulphur centred monomer radical, hydroxysulfuranyl and dimer radical of 1 respectively. The equilibrium constant between the monomer to dimer radical was 3.75?×?104 M?1. From the transients’ redox nature, it was observed that 57 and 24% of ?OH radical yielded to oxidising and reducing products respectively. Further, the product analysis by HPLC suggested that the dimer radical disproportionate to esculetin and thioesculetin. DFT energy calculation for all the possible transients revealed that dimer radical has the lowest energy. The HOMO of 1 and its monomer radical suggested that the electron density was localised on the sulphur atom. The bond length between the two sulphur atoms in dimer radical was 2.88 Å which was less than the van der Waals distance. Bond order between the two sulphur atoms was 0.55, suggesting that the bond was two centre three electron (2c–3e). From TD-DFT calculation, the electronic transition of dimer radical was at 479?nm which was in close agreement with the experimental value. The nature of the electronic transition was σ → σ* from a 2c???3e bond.  相似文献   

12.
Triosephosphate isomerase (TIM) is an enzyme with a role in glycolysis and gluconeogenesis by catalyzing the interconversion between glyceraldehyde 3-phosphate and dihydroxyacetone phosphate. This enzyme has been used as a target in endoparasite drug development. In this work we cloned, expressed, purified and studied kinetic and structural characteristics of TIM from tick embryos, Rhipicephalus (Boophilus) microplus (BmTIM). The Km and Vmax of the recombinant BmTIM with glyceraldehyde 3-phosphate as substrate, were 0.47 mM and 6031 ??mol min−1 mg protein−1, respectively. The resolution of the diffracted crystal was estimated to be 2.4 Å and the overall data showed that BmTIM is similar to other reported dimeric TIMs. However, we found that, in comparison to other TIMs, BmTIM has the highest content of cysteine residues (nine cysteine residues per monomer). Only two cysteines could make disulfide bonds in monomers of BmTIM. Furthermore, BmTIM was highly sensitive to the action of the thiol reagents dithionitrobenzoic acid and methyl methane thiosulfonate, suggesting that there are five cysteines exposed in each dimer and that these residues could be employed in the development of species-specific inhibitors.  相似文献   

13.
The c20-c21 α-ketol system of the 176 side chain of deoxycorti-costerone is in equilibrium with its c20-c21 enediol. The apparent dissociation constant of this enol was determined by a photometric method using crystal violet indicator; p Ka1 = 10.65 ± 044. Formation constants of the (1:1) deoxycorticosterone-cobalt (II) complex were determined by solvent extraction and in mixed solvent systems. The complex formation constant Kf in an aqueous medium was found by graphical extrapolation to be 2.5–3.0 × 10−1 1. mole−1  相似文献   

14.
Curcuma longa rhizome lectin, of non-seed origin having antifungal, antibacterial and α-glucosidase inhibitory activities, forms a homodimer with high thermal stability as well as acid tolerance. Size exclusion chromatography and dynamic light scattering show it to be a dimer at pH 7, but it converts to a monomer near pH 2. Circular dichroism spectra and fluorescence emission maxima are virtually indistinguishable from pH 7 to 2, indicating secondary and tertiary structures remain the same in dimer and monomer within experimental error. The tryptophan environment as probed by acrylamide quenching data yielded very similar data at pH 2 and pH 7, implying very similar folding for monomer and dimer. Differential scanning calorimetry shows a transition at 350.3 K for dimer and at 327.0 K for monomer. Thermal unfolding and chemical unfolding induced by guanidinium chloride for dimer are both reversible and can be described by two-state models. The temperatures and the denaturant concentrations at which one-half of the protein molecules are unfolded, are protein concentration-dependent for dimer but protein concentration-independent for monomer. The free energy of unfolding at 298 K was found to be 5.23 Kcal mol−1 and 14.90 Kcal mol−1 for the monomer and dimer respectively. The value of change in excess heat capacity upon protein denaturation (ΔCp) is 3.42 Kcal mol−1 K−1 for dimer. The small ΔCp for unfolding of CLA reflects a buried hydrophobic core in the folded dimeric protein. These unfolding experiments, temperature dependent circular dichroism and dynamic light scattering for the dimer at pH 7 indicate its higher stability than for the monomer at pH 2. This difference in stability of dimeric and monomeric forms highlights the contribution of inter-subunit interactions in the former.  相似文献   

15.
Plasmodium falciparum adenylosuccinate synthetase, a homodimeric enzyme, contains 10 cysteine residues per subunit. Among these, Cys250, Cys328 and Cys368 lie at the dimer interface and are not conserved across organisms. PfAdSS has a positively charged interface with the crystal structure showing additional electron density around Cys328 and Cys368. Biochemical characterization of site directed mutants followed by equilibrium unfolding studies permits elucidation of the role of interface cysteines and positively charged interface in dimer stability. Mutation of interface cysteines, Cys328 and Cys368 to serine, perturbed the monomer-dimer equilibrium in the protein with a small population of monomer being evident in the double mutant. Introduction of negative charge in the form of C328D mutation resulted in stabilization of protein dimer as evident by size exclusion chromatography at high ionic strength buffer and equilibrium unfolding in the presence of urea. These observations suggest that cysteines at the dimer interface of PfAdSS may indeed be charged and exist as thiolate anion.  相似文献   

16.
Dimethyl sulfoxide (DMSO) is commonly used as a cosolvent to improve the aqueous solubility of small organic compounds. Its use in a screen to identify novel inhibitors of the enzyme NAD(+) synthetase led to this investigation of its potential effects on the structure and stability of this 60-kD homodimeric enzyme. Although no effects are observed on the enzyme's catalytic activity, as low as 2.5% (v/v) DMSO led to demonstrable changes in the stability of the dimer and its unfolding mechanism. In the absence of DMSO, the dimer behaves hydrodynamically as a single ideal species, as determined by equilibrium analytical ultracentrifugation, and thermally unfolds according to a two-state dissociative mechanism, based on analysis by differential scanning calorimetry (DSC). In the presence of 2.5% (v/v) DMSO, an equilibrium between the dimer and monomer is now detectable with a measured dimer association constant, K(a), equal to 5.6 x 10(6)/M. DSC curve analysis is consistent with this finding. The data are best fit to a three-state sequential unfolding mechanism, most likely representing folded dimer <==> folded monomer <==> unfolded monomer. The unusually large change in the relative stabilities of dimer and monomer, e.g., the association equilibrium shifts from an infinitely large K(a) down to approximately 10(6)/M, in the presence of relatively low cosolvent concentration is surprising in view of the significant buried surface area at the dimer interface, roughly 20% of the surface area of each monomer is buried. A hypothetical structural mechanism to explain this effect is presented.  相似文献   

17.
A new assay for phosphoenolpyruvate carboxykinase has been devised whose sensitivity is sufficient to detect the formation by crude hepatic extracts of 1 × 10−11 moles of phosphoenolpyruvate. By the utilization of this assay it has been determined that the apparent Km of oxaloacetate is in the range of 1 to 5 × 10−6 M for the hepatic enzymes from rat and guinea pig. These values are approximately two orders of magnitude lower than all previous data and indicate that the enzyme is sensitive to modulations of oxaloacetate concentrations that occur physiologically.  相似文献   

18.
The fluorescence decays of pyrene in small and large unilamellar L,-dipalmitoylphosphatidylcholine vesicles have been investigated as a function of probe concentration and temperature. When the molar ratio of pyrene to phospholipid equals 1:3000, no excimer emission is observed and the fluorescence decays are mono-exponential. When this ratio is equal to or higher than 1:120, excimer formation is observed.Above the phase transition temperature the observed fluorescence decays of monomer and excimer can be adequately described by a bi-exponential function. The monomer decays can be equally well fitted to a decay law which takes into account a time-dependence in the probe diffusion rate constant. The fluorescence decay kinetics are compatible with the excimer formation scheme which is valid in an isotropic medium. The excimer lifetime and the (apparent) rate constant of excimer formation have been determined as a function of probe concentration at different temperatures above the phase transition temperature. The activation energy of excimer formation is found to be 29.4±1.3 kJ/mol. In small unilamellar vesicles the diffusion constant associated with the pyrene excimer formation process varies from 8.0x10-7 cm2/s at 40°C to 2.2x10-6 cm2/s at 70°C.Below the phase transition temperature the monomer decays can be described by a decay law which takes into account a time dependence of the rate constant of excimer formation. The lateral diffusion coefficient of pyrene calculated from the decay fitting parameters of the monomer region varies from 4.0x10-9 cm2/s at 20°C to 7.9x10-8 cm2/s at 35°C. No significant difference could be observed between the pyrene fluorescence decay kinetics in small and large unilamellar vesicles.Abbreviations SUV small unilamellar vesicles - LUV large unilamellar vesicles - DPPC dipalmitoylphosphatidylcholine - DMPC dimyristoylphosphatidylcholine - FRAP fluorescence recovery after photobleaching Part of this research has been presented at the 5th international symposium on surfactants in solution. Bordeaux, July 9th–13th 1984  相似文献   

19.
The concentration dependence of the ultraviolet absorption spectrum of aqueous solutions of caffeine has been studied. Individual species spectra have been derived for the monomer, dimer, and tetramer of caffeine. The emission spectrum of caffeine in aqueous solution and the dichroic spectra in oriented poly(vinyl alcohol) and polyethylene films have been measured. The long-wavelength tail of the absorption spectrum of caffeine in non-polar environment has been found to incorporate at least one carbonyl(π*, n) transition. Dichroic spectral data and molecular orbital calculations have been used to assign transition moment directions to the (π*,π) transitions. The lowest energy (π*,π) transition, responsible for the near-ultraviolet absorption peak in aqueous solution of caffeine, has been used for the study of degenerate exciton interactions in the dimeric species of caffeine. Assuming that the caffeine molecules in the dimer are stacked in parallel planes, theoretical calculations of the ground-state interactions and of the degenerate exciton interactions have been combined with experimental data and a unique model for the dimer of caffeine has been derived. The transfer rate of energy between the molecules in the dimer is of the order of 1013S?1.  相似文献   

20.
A single-chain Fv (scFv) fragment of anti-idiotype antibody 11-1G10, which recognizes an idiotope of anti-neuraminidase antibody NC41, was constructed by joining VH and VL domains with a (Gly4Ser)3 linker, with a pelB leader sequence, and two C-terminal FLAG tag sequences, and expressed in E. coli (10 mg/L). The 11-1G10 scFv was isolated by affinity chromatography on an anti-FLAG M2 antibody column as a 2:1 mixture of monomer and dimer forms which were separated by Superdex 75 chromatography; monomer (at 100 g/ml) was stable for 7 days at 21°C and 30 days at 4°C, whereas the dimer slowly dissociated to monomer to yield a 2:1 monomer–dimer equilibrium mixture after 30 days at 4°C. The dimer was bivalent, with each combining site binding an NC41 Fab to yield a stable complex of M r 156,000. Binding affinities, determined in solution using a BIAcore biosensor, showed that the affinity for the interaction of 11-1G10 scFv monomer with NC41 scFv monomer was five- to six-fold higher than the interaction of the parent Fab pair. This is the first example of an scFv derived from a monoclonal antibody with a higher affinity than its parent Fab.  相似文献   

设为首页 | 免责声明 | 关于勤云 | 加入收藏

Copyright©北京勤云科技发展有限公司  京ICP备09084417号